Thermomechanical modeling of slab eduction

advertisement
JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 117, B08411, doi:10.1029/2012JB009137, 2012
Thermomechanical modeling of slab eduction
T. Duretz,1,2 T. V. Gerya,2 B. J. P. Kaus,3 and T. B. Andersen4
Received 4 January 2012; revised 4 July 2012; accepted 6 July 2012; published 18 August 2012.
[1] Plate eduction is a geodynamic process characterized by normal-sense coherent motion
of previously subducted continental plate. This mechanism may occur after slab
detachment has separated the negatively buoyant oceanic plate from the positively buoyant
orogenic root. Eduction may therefore be partly responsible for exhumation of high
pressure rocks and late orogenic extension. We used two-dimensional thermomechanical
modeling to investigate the main features of the plate eduction model. The results show that
eduction can lead to the quasi adiabatic decompression of the subducted crust (≈2 GPa) in a
timespan of 5 My, large localized extensional strain in the former subduction channel,
flattening of the slab, and a topographic uplift associated with extension of the orogen. In
order to further investigate the forces involved in the eduction process, we ran systematic
parametric simulations and compared them to analytic plate velocity estimations. These
experiments showed that eduction is a plausible mechanism as long as the viscosity of the
asthenospheric mantle is lower than 1022 Pa.s while subduction channel viscosity does not
exceed 1021 Pa.s. We suggest that eduction can be a viable geodynamic mechanism and
discuss its potential role during the orogenic evolution of the Norwegian Caledonides.
Citation: Duretz, T., T. V. Gerya, B. J. P. Kaus, and T. B. Andersen (2012), Thermomechanical modeling of slab eduction,
J. Geophys. Res., 117, B08411, doi:10.1029/2012JB009137.
1. Introduction
1.1. Background
[2] The exhumation of high pressure and ultra high pressure (HP-UHP) metamorphic rocks at convergent margins is
a very active field of research in tectonics and geodynamics.
Understanding the long term dynamics of subductioncollision zones requires conceptual models, which can account
for both real Earth geological and geophysical observations
and comply with quantitative geomechanical models.
[3] Numerous models specifically focused on the processes
driving the exhumation of HP-UHP rocks are available in the
literature. Syn-to-late collisional exhumation models involve
mechanisms such as corner flow or buoyant flow within the
subduction channel [Cloos and Shreve, 1988; Gerya et al.,
2002; Yamato et al., 2008; Li and Gerya, 2009], buoyancy
driven crustal stacking and development of normal sense shear
zones [Chemenda et al., 1995, 1996], crustal flow associated
with the gravitational spreading of orogens [Andersen and
Jamtveit, 1990; Vanderhaeghe and Teyssier, 2001] and focused
1
ISTEP, UMR 7193, UPMC Paris 06, CNRS, Paris, France.
Department of Earth Sciences, Institute of Geophysics, ETH Zürich,
Zürich, Switzerland.
3
Institute of Geosciences, Johannes Gutenberg University Mainz,
Mainz, Germany.
4
Physics of Geological Processes, University of Oslo, Oslo, Norway.
2
Corresponding author: T. Duretz, Department of Earth Sciences,
Institute of Geophysics, ETH Zürich, Sonneggstr. 5, Zürich CH-8092,
Switzerland. (Thibault.Duretz@erdw.ethz.ch)
©2012. American Geophysical Union. All Rights Reserved.
0148-0227/12/2012JB009137
erosion [Beaumont et al., 2001, 2004, 2006], slab rollback
driven by the retreat of the subducting slab [Lister et al.,
2001; Schellart et al., 2006; Brun and Facenna, 2008;
Husson et al., 2009; Bialas et al., 2011] and delamination of
the crust [Bird, 1978], or slab extraction involving slab
detachment and the decompression of buried material by
plate unbending [Froitzheim et al., 2003; Janák et al., 2006].
[4] The slab detachment model necessitates the build up
tensional stresses that overcomes the strength of the subducting plate. Such scenario is likely to take place during an
attempted ridge subduction or a continental crust subduction/
collision. Both of these contexts have been subject to extensive
two-dimensional studies [Davies and von Blanckenburg, 1995;
Buiter et al., 2002; Andrews and Billen, 2009; Baumann et al.,
2009; Duretz et al., 2011] as well as three-dimensional modeling [Burkett and Billen, 2011; van Hunen and Allen, 2011]
and analytical studies focused on necking dynamics
[Schmalholz, 2011]. All models agree on the fact that slab
detachment causes a dramatic change in the orogenic force
balance. Yet, the dynamic consequences of this force balance
perturbation have not yet been studied in detail.
1.2. The Term “Eduction”
[5] The concept of plate eduction has been introduced by
Dixon and Farrar [1980] to describe a mechanism that
could lead to the exhumation of subducted/accreted rocks at
an ocean-continent margin. Their model, conceived for the
exhumation of the Californian Franciscan blueschists,
involved the subduction of an actively spreading ridge
beneath North America. The subduction of a ridge triggers
shallowing of the slab and the ongoing spreading promotes
extension in the subducting slab which would eventually
B08411
1 of 17
B08411
B08411
DURETZ ET AL.: SLAB EDUCTION
Figure 1. Initial distribution of the compositional fields employed in the continental collision experiments.
The initial thermal age of the slab is 40 My. The arrows indicate the location where the push condition is
initially applied. A total convergence rate of 5 cm/y is imposed until 500 km of convergence has been
reached.
lead to the exhumation of the subducted/accreted material
along the margin.
[6] Andersen et al. [1991] introduced the subductioneduction model to explain burial and exhumation of HP-UHP
rocks in continent-continent collision. Based on geological
data from one the largest and best preserved HP-UHP provinces in the world this work highlighted the potential link
between slab detachment, orogenic extension and exhumation of coherent slab of HP-UHP rocks through the evolution
of the Norwegian Caledonides. In this model, the continental
lithosphere was subducted to the point at which slab
detachment occurred, causing the removal of the slab pull
force. Subsequently, the subducted and vertically stretched
continental plate was coherently educted leading to the
exhumation of HP-UHP rocks along a large normal-sense
shear zone near the former subduction plane. This eduction
concept has remained a popular model in the literature
focused on postorogenic extension, specifically for the case of
the Caledonides [Fossen, 2000; Brueckner and van Roermund,
2004; Brueckner, 2009; Rey et al., 1997; Schlindwein and
Jokat, 2000] and the Variscides [Schneider et al., 2006]. This
definition contrast with that of [Dixon and Farrar, 1980] since
it explicitly implies a period of reversed subduction following
slab detachment during continental collision.
1.3. Present Work
[7] In this paper we concentrate on the large scale geodynamic process of lower plate eduction in a continental
collisional context (i.e. definition of Andersen et al. [1991]).
Post slab detachment eduction was observed in several previous numerical modeling studies [Mishin et al., 2008;
Duretz et al., 2011] but has not yet been subjected to a
systematic parametric study. We therefore focus on the
description of the major features of the eduction model and
study its dynamics by means of two-dimensional numerical
modeling. Finally, we discuss the potential role of eduction
for the exhumation of HP-UHP rocks and focus on the case
of the Norwegian Caledonides.
2. Numerical Modeling
2.1. Setup
[8] We have used a setup consisting of two continents and
of, initially, one ocean (Figure 1). The simulations were run
with the thermomechanical code I2VIS [Gerya and Yuen,
2003a], a description of the code is provided in Appendix A.
Each lithology was characterized by a temperature-stress
dependent visco-plastic rheology, with rheological and flow
parameters as listed in Table 1. Following the approach of
Gerya et al. [2004], all simulations included the effect of
phases changes on material densities neglecting the effects
of reaction kinetics. The size of the model domain was
4000 ! 1400 km and variable grid spacing (1361 ! 351 nodes)
was employed to attain a 1 km grid spacing in the collision
area. We initially imposed a plate convergence rate of 5 cm/y
by prescribing velocities inside the domain. As soon as the
Table 1. Thermal and Rheological Parameters for the Lithologies Employed in the Simulationsa
Material
Sediments
Upper cont. crust
Lower cont. crust
Upper oceanic crust
Lower oceanic crust
Mantle
Weak zone
k (W/m/K)
0.64 +
0.64 +
1.18 +
0.64 +
1.18 +
0.73 +
0.73 +
807
Tþ77
807
Tþ77
474
Tþ77
807
Tþ77
474
Tþ77
1293
Tþ77
1293
Tþ77
Hr (W/m3)
1.50 !
1.00 !
0.25 !
0.25 !
0.25 !
2.20 !
2.20 !
#6
10
10#6
10#6
10#6
10#6
10#8
10#8
Cp (J/kg)
1000
1000
1000
1000
1000
1000
1000
Flow Law
wet Qz.
wet Qz.
Pl. (An75)
wet Qz.
Pl. (An75)
dry Ol.
wet Ol.
h0 (Pan.s)
17
1.97 ! 10
1.97 ! 1017
4.80 ! 1022
1.97 ! 1017
4.80 ! 1022
3.98 ! 1016
5.01 ! 1020
n
2.3
2.3
3.2
2.3
3.2
3.5
4.0
Ea (J)
5
1.54 ! 10
1.54 ! 105
2.38 ! 105
1.54 ! 105
2.38 ! 105
5.32 ! 105
4.70 ! 105
Va (J/bar)
sin(f)
C (MPa)
0.8
0.8
1.2
0.8
0.8
0.8
0.8
0.15
0.15
0.15
0.00
0.60
0.60
0.00
1
1
1
1
1
1
1
a
Here k is the thermal conductivity, Hr is the radiogenic heat production, Cp is the specific heat capacity, h0 is the reference viscosity, n is the stress
exponent, Ea is the activation energy, Va is the activation volume, f is the internal friction angle, and C is the cohesion. Qz., Pl., and Ol. respectively
stands for Quartzite, Plagioclase, and Olivine.
2 of 17
B08411
DURETZ ET AL.: SLAB EDUCTION
B08411
Figure 2. Large scale time evolution of the reference subduction/collision model. The two plates converge toward each other leaving space for the development of ridges in the vicinity of the model’s left/right
boundaries. The interaction of the detached slab with the 660 km phase boundary is also depicted.
model reached 500 km of convergence, the kinematic condition was deactivated and the models were subsequently
driven by internal forces (e.g. slab pull). All the boundaries of
the box were free slip. An additional, 20 km thick layer of
sticky air (rair = 0 kg/m3, hair = 1018 Pa.s) was utilized in
order to mimic the effect of a free surface and the development
of topography [Schmeling et al., 2008; Crameri et al., 2012].
The setup was designed such that the continental plates detach
from the sides of the box during convergence, leaving space
for the development of oceanic ridges in the vicinity of the
domain boundaries (see Figure 2). In the simulations presented
here, the initial continental crustal thickness was 35 km. We do
not vary this parameter in this study, nevertheless we expect
that variations in crustal thickness (and therefore of plate
buoyancy) will affect both the burial depth of the continental
crust and the amount of subsequent eduction that can be
accommodated.
lithosphere may exhume coherently due its buoyancy triggering plate eduction. As presented in Duretz et al. [2011], plate
eduction is likely to occur for continental collision involving
relatively fast plate velocities (5–10 cm/y) and initial slab
thermal ages ranging between 40 and 60 My. Younger slabs
are too weak and detach before any continental crust subduction happens. Older slabs lead to a collisional regime dominated by retreat and delamination in the sense of Bird [1978].
It is important to notice that, in these models, eduction
develops because no convergent kinematic constrains are
imposed during the collision. Although the plates are laterally confined by newly formed oceanic lithosphere, the
buoyancy of the root may be sufficiently large to overcome
the ridge push force leading to divergent plate motion and
inversion of the subduction plane.
2.2. Potential Causes for Eduction
[9] We assume that the closure of an oceanic basin occurs
prior to continental collision and that this closure stage is
accommodated by oceanic plate subduction. Depending on
the dimensions, convergence rate and density structure of
the oceanic basin, subduction generates the slab pull force
which is generally considered as a major tectonic force
[Turcotte and Schubert, 1982]. While continental margins
enter into contact, a period of continental subduction can
occur. Due to the buoyant nature of the crust, continental
subduction is not steady and might terminate by slab detachment. Once slab detachment has occurred, and assuming a
competent crustal rheology, the subducted continental
3.1. Time Evolution
[10] The model evolution can be decomposed in several
successive stages (Figures 2 and 3). The earliest stage is the
closure of the oceanic basin. During this period the slab pull
(FP) builds up progressively as oceanic plate subduction
proceeds (Figure 4b). After 11 My, continental subduction
initiates and continues until the ocean-continent transition
(OCT) reaches a depth at which the positive buoyancy of the
crust (FB) compensates the negative buoyancy of the slab
(FP ≈ FB ≈ 3 ! 1013 N/m). This period lasts approximately
6.5 My and the OCT is buried to a maximum depth of
170 km. Slab detachment eventually occurs and necking
takes place at the OCT separating the negative (oceanic) from
3. The Plate Eduction Model: Reference Model
3 of 17
B08411
DURETZ ET AL.: SLAB EDUCTION
B08411
Figure 3. Time evolution of the plate eduction model in six stages focused on the collision zone.
The white arrows highlight the sense of shear on the subduction plane, the stretching direction of the slab
during detachment and the direction of flow during late exhumation. In this figure, the time is incremented
from the start of the experiment. The orange/brown layers represents sediments that are deposited with
ongoing subsidence.
the positively buoyant (continental) portions of the slab.
Figure 4c show the average density difference between the
subducted continental lithosphere and the surrounding
asthenosphere. This density contrast is responsible for the
build up of buoyant stress within the subducted crust and the
large magnitude of the root buoyancy force. After the oceanic
slab has detached, the slab pull force decreases dramatically
(Figure 4b) and the system is driven by the continental root
buoyancy and the ridge push (≈1 # 5 ! 1012 N/m in our
models). As a result of the imbalance between these two
forces, the subduction plane reverses as a normal-sense shear
zone along which the positive buoyancy of the subducted
crust is accommodated. Throughout the eduction stage that
lasts 5 My, the main portion of the slab returns toward the
surface in a coherent motion. After a progressive decrease of
the root buoyancy force (Figure 4b), the orogen is eventually
affected by a period of buoyant flow which causes the continental crust to rise diapirically toward the surface of the
model (see Figure 3).
3.2. Exhumation of the Subducted Crust
[11] In this model the exhumation occurs by two different
mechanisms: plate eduction and buoyant flow. Eduction
allows for a vertical displacement of the tip of the slab of
about 60 km. Buoyant flow of the crust overtakes eduction at
its latest stage and is responsible for doming at shallow
structural level. These two exhumation stages are clearly
identifiable on the pressure-temperature-time (PTt) paths of
material tracers (Figure 5). These clockwise paths display a
pressure peak corresponding to timing of slab detachment
followed by a nearly adiabatic decompression corresponding
to the eduction stage. The later cooling event occurs when
buoyant flow takes over eduction, this process occurs at a
crustal level (depth < 35 km). The signal of slab detachment,
and subsequent eduction, can also be noticed on the vertical
velocity (vy) evolution of the material tracers (Figure 5).
Slab detachment is followed by an “instantaneous” peak
exhumation rate of 8 cm/y that exponentially decreases to a
rate of about 0.8 cm/y through the period of eduction.
3.3. Topographic Evolution
[12] Plate eduction is responsible for a major uplift event
or rebound following slab detachment. The topographic
evolution of the subduction-collision zone can be divided in
four geodynamic phases that are characterized by a specific
surface imprint (Figure 6a). During the first 11 My period,
closure of the oceanic basin takes place, collision takes
over oceanic subduction and lasts until 18 My when slab
4 of 17
B08411
DURETZ ET AL.: SLAB EDUCTION
B08411
Figure 4. (a) Post slab detachment density structure. The white contours denotes lithological groups
(sediments, upper crust, lower crust, mantle lithosphere and asthenosphere). (b) Time evolution of slab
pull and root buoyancy forces (buried continental crust) throughout the subduction/collision history.
The time is incremented from the onset of the experiment. (c) Average vertical profile of density difference
between the subducted continental lithosphere and the asthenosphere after slab detachment. This average
density contrast (Dr) is computed by first evaluating mean mantle and crust densities at each depth level
(i.e. horizontal gridlines), and by subtracted one from the other.
detachment happens. While continental subduction proceeds,
the topographic peak is located on the upper plate whereas a
deep subduction trench is located on the lower plate. During
the next 6 My following slab detachment, the topography
builds up on both sides of the suture marking the stage of
plate eduction. The uplift also takes place in the neighboring
foreland and hinterland basins. The smooth transition
between eduction and buoyant flow that occurs at around
5 of 17
B08411
DURETZ ET AL.: SLAB EDUCTION
B08411
Figure 5. Pressure, temperature, density, vertical velocity evolution for 4 Lagrangian material tracers
located in the subducted crust. The left column represents the location of the markers throughout the exhumation period. Dashed lines indicate the signal corresponding to slab detachment/start of eduction and the
end of the eduction/start of buoyant flow. The time is incremented from the onset of the collision.
25 My of model time does not result in a sharp topographic
feature.
3.4. Plate Deformation and Motion
[13] Eduction is characterized by a rigid-body (or coherent)
motion of the lower plate. The transition from a subductionrelated plate motion to eduction is sharp and is depicted in
Figure 6b. This divergent motion causes localized extension
within the subduction channel. The finite deformation pattern
(Figure 7), produced using the methodology of Huet et al.
[2010], reveals the intensity of the strain that accumulated
along the subduction channel and the asthenospheric mantle
underlying the plate. This illustrates the coherent plate
motion resulting from the slab pull loss and the absence of
kinematic constrains. The educted plate itself does not
undergo any significant internal deformation apart from the
pervasive but not intense strain resulting from the unbending.
The surface plate velocities also highlight the rigid motion of
lower plate throughout the collision. After slab detachment,
the lower plate reaches a velocity on the order of #5 cm/y
during few My which coincides with eduction (Figure 6b).
3.5. Slab Unbending/Flattening
[14] Another feature of the eduction model is tendency for
the lower plate to unbend throughout the eduction stage.
Figure 6c depicts the evolution of the slab dip angle after
slab detachment occurred, the slab dip is here defined as the
angle between the interface separating the underthrusted
continental crust and the subduction channel. The normal
shear motion along the former subduction interface is
accompanied by the unbending of the lower plate and the
flattening of the slab. In our reference model, the dip of the
slab decreases from 35$ to about 22$ through the extensional
period (5 My timespan).
4. The Dynamics of Plate Eduction
[15] As described above, plate eduction is a dynamic
consequence of the loss of slab pull subsequent to slab
detachment. In order to investigate which parameters control
the post slab detachment tectonic regime, we have combined
both an analytic approach (Figure 8a) and a simplified 2D
numerical model (Figure 8b). The analytic method enables
6 of 17
B08411
B08411
DURETZ ET AL.: SLAB EDUCTION
Figure 6. (a) Topographic evolution through time of the reference eduction model, white dashed lines
indicate major geodynamic events affecting the topography. The two profiles on the right hand side
display the topography along the collision zones before and after eduction. The time is incremented from
the start of the experiment. (b) Horizontal plate velocity during continental subduction, slab detachment,
and eduction. (c) Slab dip evolution through the period of eduction following slab detachment. The time
axis origin is the timing of slab detachment.
to test the influence of the force difference acting on a plate
whereas the numerical model allows for the independent
variation of the orogenic root buoyancy and the ridge push
force. These calculations are utilized to discriminate under
which force regime subduction or eduction occurs.
5. A Corner Flow/Torque Balance Approach
[16] In order to obtain a first order prediction of subduction velocity we have used the classical torque balance
approach [Stevenson and Turner, 1977; Manea et al., 2006].
This approach is isoviscous and allows us to investigate the
dependence of subduction velocity on the mantle viscosity
and the slab angle. The torque balance formulates the equilibrium between the moments produced by gravity MG (i.e.
slab’s buoyancy) and mantle flow MF (i.e. “slab’s lifting
torque”). This simplistic approach neglects the torque related
to slab bending [Dvorkin et al., 1993; Lallemand et al.,
2008].
[17] The hydrodynamic torque resulting from the arc and
back arc pressures difference (PA # PB) is obtained from
the isoviscous stream function approach [McKenzie, 1969]
expressed in polar co-ordinates (r, q) and can be formulated
as follows:
MF ¼
Z
L
0
ðPA # PB Þrdr:
ð1Þ
[18] The arc/back arc pressure difference is expressed as in
[Stevenson and Turner, 1977]:
7 of 17
"
#
2hast V
sin qs
sin2 qs
PA # PB ¼
þ
;
ðp # qs Þ þ sin qs q2s # sin2 qs
r
ð2Þ
B08411
B08411
DURETZ ET AL.: SLAB EDUCTION
Figure 7. Finite strain intensity pattern related to eduction. Most of the deformation is accommodated by
the subduction channel and the asthenosphere. The strain is incremented after the slab has detached. The
time after collision has started is indicated. The white contours represent the interfaces between lithological groups (sediments, upper crust, lower crust, mantle lithosphere and asthenosphere).
where qs is the slab dip, hast is the asthenosphere viscosity,
and V is the slab velocity. The gravitational torque results
from the density difference existing between the slab and
mantle (Dr):
MG ¼
Z
L
0
Drgh cos qs rdr;
ð3Þ
where g and h respectively stands for the gravitational
acceleration and the slab thickness. Integrating and solving
both expressions for V leads to the velocity expression:
V ¼
F K
;
4 hast
where F = DrghL is the slab’s buoyancy force and K is a
factor that depends on the slab dip:
K¼
cos qs
sin qs
ðp#qs Þþ sin qs
2
þ q2 #sinsinq2s q
s
:
ð5Þ
s
[19] Assuming that the slab is rigid and transmits the farfield ridge push force, we may consider that the slab is
effectively driven by the force difference between slab’s
buoyancy and ridge push acting in the direction of the slab
(F ∝ dF = FP cos qs # FB sin qs). Subsequently, the velocity
expression can be rewritten as
ð4Þ
8 of 17
V ∝
dF K
;
4 hast
ð6Þ
B08411
B08411
DURETZ ET AL.: SLAB EDUCTION
Figure 8. (a) Sketch of the corner flow/torque balance model used to calculated plate velocities. The plate
is considered as rigid and the velocities are calculated from the balance between the gravitation torque
and the hydrodynamic torque obtained from the corner flow model. (b) Simplified setup of a post slab
detachment configuration. This setup is used to systematically investigate the parameters that control
horizontal plate velocity (subduction or eduction) by varying the ridge push and the orogenic root buoyancy
as well as the asthenosphere and subduction channel viscosity, the lower plate length (L) and the subduction
angle (q).
where the velocity of plate motion is directly proportional to
dF and which is either positive when subduction occurs or
negative when eduction takes place.
where rlit, rast2, and Aridge respectively stand for the lithosphere density, the ridge and the ridge area. Similarly, we
define the magnitude of the orogenic root buoyancy as:
FB ¼ ðrlit # rroot ÞAroot g;
6. Simplified 2D Setup
[20] In order to test the applicability of the torque balance
approach, we performed additional numerical simulations
for a simplified setup that consists of a rigid newtonian
(high-viscosity) slab and a linear viscous mantle. The steady
state equation of momentum and incompressibility are discretized using T2P1 elements and solved for the given distribution of density and viscosity, after which the resulting
instantaneous flow pattern is used to determine whether the
orogen is undergoing subduction (Vplate > 0) or eduction
(Vplate < 0). The boundary conditions are set to free slip on
the left, right and bottom boundaries, and a free surface
condition is applied at the top of the box. All the simulations were run using the code MILAMIN_EP [Kaus, 2010].
An unstructured mesh was employed to finely resolve the
subduction interface. This method allows for fast and robust
computation of flow fields as required for parametric studies.
[21] The horizontal (i.e. ridge push) force is prescribed by
varying the density gradient between the asthenosphere and
the plate, whereas the root buoyancy force is controlled by
the density difference between the buried crust and the surrounding mantle (Figure 8b). We define the magnitude of
the ridge push force as
FP ¼ ðrlit # rridge ÞAridge g;
ð7Þ
ð8Þ
where rroot and Aroot correspond to the density and the area
of buried continental crust and are calculated according to
the magnitude of each force such as:
rridge ¼ rlit #
FP
Aridge g
ð9Þ
rroot ¼ rlit #
FB
:
Aroot g
ð10Þ
The tectonic regime is therefore controlled by the forces that
are applied to the plates and imposed via the density distribution. In contrast to the isoviscous torque balance described
in 5, such setup enables to explore the effect of the subduction channel rheology on plate kinematics.
6.1. Parameters Controlling Eduction
[22] Over the many parameters that affect the velocity of
the subducting or educting plates, we have focused on the
dip angle of slab and the viscosities of both the asthenosphere and the subduction channel. Both torque balance and
2D finite element (FE) models were employed to calculate
plates velocity after slab detachment (the results are depicted
in Figure 9a). Despite the different assumptions made with
9 of 17
B08411
DURETZ ET AL.: SLAB EDUCTION
B08411
Figure 9. (a) Plate velocity for variable force difference (dF) (using hast = 1020 Pa.s and q = 30$ ).
The solid lines represent the solution obtained from the 2D simulations for different subduction channel
viscosities (hch). The dotted line is the prediction of the torque balance approach for an isoviscous mantle.
(b) Sensitivity of plate velocity to the slab dip angle for variable force difference values. This was computed using the torque balance model. (c) Magnitude of the basal drag (red) and channel drag forces (blue)
estimated for our 2D setup as a function of plate velocity. These estimations assume a Couette flow in the
asthenosphere and the subduction channel. (d) Efficiency of the subduction channel drag over the basal
drag for various channel aspect ratios and channel/asthenosphere viscosity contrast.
each approach, the FE and torque balance results are in good
agreement. The FE models confirm the result of the toque
balance approach when the subduction channel’s viscosity
is equal to that of the asthenosphere (isoviscous case).
Both approaches predict that eduction occurs if the buoyancy
of the slab exceeds the ridge push force (dF < 0).
6.1.1. Slab Dip
[23] The effect of the slab dip was investigated using the
torque balance approach. We have used a constant
asthenosphere viscosity of 1020 Pa.s and varied the slab dip
from 10 to 70 degrees (Figure 9b). The slab dip angle shows
a significant influence on the plate velocities. Plate velocities
increase with increasing dip angle up to an optimal angle of
≈60 degrees (close to the results of Stevenson and Turner
[1977]). Over this critical dip angle value, the plates velocities decrease with increasing slab dip.
6.1.2. Asthenosphere Viscosity
[24] The asthenosphere provides viscous resistance to slab
penetration. This resistance can be decomposed in force
acting parallel to the motion direction and a basal drag tangential to the slab face (FBD). From the torque balance model,
the slab velocity is inversely proportional to the asthenosphere viscosity. A viscosity increase of one order of magnitude will therefore produce a velocity decrease of the same
magnitude. However, a more realistic model would also take
into account effects of the slab bending [Buffet, 2006], the
drag torque due to slab curvature, or the slab anchoring
due to the upper plate motion [Scholz and Campos, 1995;
10 of 17
B08411
DURETZ ET AL.: SLAB EDUCTION
Lallemand et al., 2008] which would also affect subduction/
eduction rates. The magnitude of the basal drag force in or
2D can be estimated by considering a Couette flow in the
mantle driven by the slab motion in such manner:
FBD ∝ hast V
LPlate
:
Hast
ð11Þ
Using a plate length (LPlate) of 1500 km and an asthenospheric channel thickness (Hast) of 500 km, Figure 9c shows
that the estimated magnitude of basal drag does not exceed
8 ! 1011 N/m even for plate velocities larger than 8 cm/y.
This justifies that the basal drag force plays a minor role in
our calculations.
6.1.3. Subduction Channel Viscosity
[25] Similarly to the basal drag force, a slab tangential
force is also generated in the subduction channel (FBD). Its
influence on plate velocities can not be investigated by the
isoviscous torque balance model, however the 2D simulations showed large sensitivity to this parameter (Figure 9a).
At first order, the magnitude of this force may also be
approximated from a simple Couette flow model driven by
the subducting plate and leading to the expression:
FCD ∝ hch V
LChannel
:
HChannel
ð12Þ
Estimated channel drag forces computed using a channel
length (LChannel = 300 km) and thickness (HChannel = 15 km)
corresponding to our 2D setup are depicted in Figure 9c.
In the ideal case, the channel is considered to provide strong
mechanical decoupling between the plates, and hence should
not produce a strong drag force (e.g. Figure 9c with
hch = 1018 Pa.s). However, its magnitude is highly dependent
on the channel viscosity and may overcome the basal drag
force once the channel viscosity tends to the asthenosphere
viscosity (e.g. Figure 9c with hch = 1020 Pa.s). Another
major parameter is the aspect ratio of the channel, Figure 9d
shows the ratio of channel/basal drag force magnitudes for
variable channel/asthenosphere viscosity ratio and channel
aspect ratio. These results indicate that, in the case the channel
undergoes strengthening processes and/or geometrical modifications (thickening/thinning), channel drag may largely
overcome basal drag and might become a major force
(FCH > 1012 N/m).
7. Discussion
7.1. Low Peak Temperature in the Exhumed Crust
[26] In our simulations, the reference model presented here
fails at producing sufficiently large enough temperatures in
the exhumed HP rocks compared to geological examples of
HP and UHP rocks formed at corresponding extreme burial
depths (T < 600 to 750$ C). These low temperatures can be
explained by the short residence time of subducted material
at mantle depths which is limited by the onset of extensional
processes within the orogen. An unrealistic low geotherm of
less than 3 C/km during continental subduction is the result
of this model. Hence, a fast (cold) subduction/collision is
subject to a rapid burial and aborted by slab detachment.
In this case, the main exhumation stage occurs after the
detachment and the buried continental margin is educted
B08411
without reaching high temperatures. Since a fast convergence plate eduction models can only explain the exhumation of cold UHP rocks (T < 500$ C), we expect that slower
convergence rates are likely to produce and exhume UHP
rocks with higher temperatures. Although our simulations
take into account the thermomechanical feedback induced
by viscous dissipation, the peak temperatures reached by the
exhumed material are approximately 100$ C too cold to
match natural data. These results, comparable to those
obtained in previous studies [Yamato et al., 2008; Stöckhert
and Gerya, 2005], tend to show that viscous dissipation may
not explain common peak temperatures (T > 600$ C) of
exhumed HP and UHP rocks. In contrast to the results of
Li and Gerya [2009] and Li et al. [2011], our simulations do
not account for partial melting, this may potentially explain
our low peak temperatures. We expect the heat advected by
the partially molten upwellings and the heat released during
crystallization of potential plutons to provide the missing
≈100$ C. As material for discussion, we present additional
subduction-collision models that were obtained following
the same methodology as in the simulations presented above
(i.e. without including any effect related to melting).
7.1.1. A Delamination-Eduction Model
[27] The models presented in Section 3 were designed in
order to isolate the process of slab eduction. However, more
realistic models of orogenic evolution may include delamination [Bird, 1978] or buoyant uplift [Warren et al., 2008a,
2008b] linked to slab rollback. Figure 10a depicts the evolution of an initially slower continental collision (1.25 cm/y).
The simulation runs for 40 My before continental subduction
takes place. The burial stage lasts for about 10 My and
the subducted crust is affected by delamination as soon as
the continental crust reaches its maximum depth of burial
(≈200 km). Most of the decompression occurs during the
delamination period that lasts approximately 7 My before
slab detachment eventually occurs triggering a late stage
eduction event. As indicated by Figure 10b, the exhumed
crust peak temperatures higher than in the case of pure
eduction (≈660$ C). Another feature of this delaminationeduction model is the flow of the extruded crust toward the
foreland that leads to a stage of orogenic broadening, and the
occurrence of a deep-seated thrust that likely to be exposed in
the hinterland of the broadened mountain belt. The combination of delamination and eduction gives rise to the exhumation of HP-UHP rocks with peak temperatures of 500 to
650$ C, more akin to natural examples, and we therefore
regard this type of exhumation model as more realistic.
7.1.2. Subduction of a Thinned Continental Margin
[28] Similarly to boudinage instabilities [Schmalholz et al.,
2008], slab detachment has been demonstrated to be the
result of a viscous necking instability in a power law fluid
[Schmalholz, 2011]. We therefore expect our slab detachment to be sensitive to the geometry and the thermal state of
the subducted margin (amplitude and wavelength of the
initial perturbation). Figure 11a displays the time evolution
of a model employing a thinned ocean-continent transition.
Although, the initial convergence rate is prescribed similarly as for our reference model (5 cm/y), the presence of a
stretched ocean-continent transition delays slab detachment
and ultimately favors delamination. The distal part of the
margin reaches a maximum depth larger than 200 km from
where the delamination of the upper crust is initiated. As the
11 of 17
B08411
DURETZ ET AL.: SLAB EDUCTION
Figure 10. (a) Temporal evolution of the delamination/eduction model, the colored stars correspond to
Lagrangian tracers used to record the pressure-temperature history of the subducted continental crust.
(b) Pressure-temperature evolution of the Lagrangian tracers throughout the collision event.
Figure 11. (a) Evolution of the thinned margin subduction model, this feature inhibits slab detachment at
the timescale of the collision. The Lagrangian tracers used to record the pressure-temperature history of
the subducted continental crust are denoted as colored stars. (b) PT paths corresponding to different initial
locations within the passive margin.
12 of 17
B08411
B08411
DURETZ ET AL.: SLAB EDUCTION
delamination occurs (≈10 My), several nappes detach from
the subducted margin. These buoyant nappes are emplaced
adjacently stacked within the orogen and record significantly different PT histories (Figure 11b). This model is in
agreement with previous studies [Warren et al., 2008b; Li
et al., 2011; Yamato et al., 2008] that showed that slab
detachment is not a necessary factor for the exhumation of
high pressure rocks and that the stage of continental collision
does not exceed more than 10 My [Yamato et al., 2008].
7.1.3. The Western Gneiss Region, Scandinavian
Caledonides
[29] The idea of continental subduction and eduction following slab breakoff at the terminal stages of a Wilson cycle
was put forward in order to explain the geological observations from the Scandian continent-continent collision in the
Caledonides [Andersen et al., 1991]. The collision produced
one of the worlds largest HP-UHP terrains and the main
characteristics of the vast Western Gneiss Region (WGR)
are:
[30] 1. Protoliths of the HP-UHP rocks are Middle Proterozoic (≈1700 to 950 My) mostly orthogneisses [e.g.,
Austrheim et al., 2003; Tucker et al., 2004], and major parts
of the WGR experienced granulite metamorphic conditions
at ≈1 Ga [e.g., Røhr et al., 2004; Krabbendam et al., 2000].
[31] 2. The uppermost structural level of the WGR is a
nearly intact Proterozoic crust, only affected by the Caledonian metamorphism and deformation near the basal Caledonian thrust [Labrousse et al., 2010].
[32] 3. The WGR has a metamorphic field-gradient from
600$ C at 1.8 GPa in the east to 750$ C at 2.8 GPa in the
west [e.g., Young et al., 2007; Hacker et al., 2010].
[33] 4. The HP-UHP metamorphism in the WGR lasted
nearly 20 My (415 to 397 My [e.g., Krogh et al., 2011]),
whereas Scandian HP in the nappes is older up to 430 My
[Glodny et al., 2008].
[34] 5. There are no exposed major syn- to post-UHPmetamorphic thrusts [Hacker et al., 2010], and a very largescale extensional detachment zone separates the HP-UHP
rocks in the footwall from lower grade nappes in the hanging
wall.
[35] 6. The total duration of the Scandian continental collision was approximately 30 My (430 to 400 My) [Andersen
et al., 1990; Corfu et al., 2006].
[36] The absence of large-scale structures perturbing the
regional metamorphic zonation within the WGR suggests
that it was buried and exhumed as a mostly intact slab
(40000 km2) of continental crust. Some very local extreme
UHP occurrences recording 3 to 6 GPa (diamond, majoritic
garnet and opx-eclogites) within the coesite-grade domains
(see review by van Roermund [2009]) cannot be fitted to the
regional metamorphic field gradient and probably needs
alternative explanations [Vrijmoed et al., 2009], which will
not be discussed here. The eduction model presented in this
study (Figure 5) results in a 2D structural geometry of the
exhumed HP-UHP rocks, which is quite similar to observations in the WGR [Andersen et al., 1991, Figure 2]. The
duration since on-set of collision is 27 My (Figure 4b), also
similar to the Caledonian analog. The HP-UHP rocks also
constitute the lowermost observable structural level similar
to what is observed in western Norway. There is no largescale thrust below the UHP rocks, and there is a very large
B08411
extensional detachment above the HP-UHP rocks. The model
presented in Figure 5, however, fails to reproduce the
temperatures similar to those seen in the eclogites in the
WGR by approximately 150 to 250$ C [Hacker et al.,
2010; Labrousse et al., 2004]. It is obvious that the reference eduction model used here is too cold to explain the
WGR example in full. The alternative delamination-eduction
model (Figure 10) also reproduces a number of the field
characteristics, and the temperatures obtained are still lower
(up to 100$ C), but more comparable to those from the WGR.
The main structural difference between the eduction and the
delamination-eduction models is that the delamination stage
produces a major thrust at mid-crustal levels, below the UHP
rocks during exhumation (Figure 10). The presence of such
thrust(s) has been postulated in several papers on the WGR
[e.g., Andersen et al., 1991; Hacker et al., 2010], but the
existence has not been verified by observation in the field.
If present it must be covered by nappes in south-central
Norway. Observations that argue against the presence of the
deep thrust is that sedimentary cover to the autochthonous
Proterozoic basement can be traced almost continuously
along the basal thrust from the foreland below the nappes to
the WGR in central south Norway. Considering all available
geological and geophysical observations [Andersen, 1998],
a geomechanical model combining delamination with slab
breakoff and eduction provides the best quantifiable model
for exhumation of the HP-UHP rocks in the WGR.
7.2. Limitations of the Density Model
[37] Although we did not explicitly consider the effect of
phase transitions occurring within the continental crust, the
crustal densities were calculated according to the transient
pressure-temperature conditions. The equation of state (see
Appendix A) allows the upper crustal density to smoothly
vary between 2700 to 2850 kg/m3 whereas lower crust
density ranges between 2800 to 2950 kg/m3 (Figure 4).
Both density changes related to either phase changes or the
equation of state were considered to be instantaneous and
reversible. Natural observation indicates that reaction kinetics
can be slow and can lead to partial prograde phase transformations [Krabbendam et al., 2000] and neglecting this effect
might lead to an overestimation of the crust density on the
prograde path. This would have the effect of overvaluing the
buoyancy of the crust and could potentially reduce the peak
burial depth of continental crust prior to slab detachment.
On the other hand, retrograde density changes may lead to
an underestimation of the crust density and therefore overrating the root buoyancy force in the later stages of eduction.
7.3. Comparison With the Models of Coherent
Nappe Exhumation
[38] The models proposed by Chemenda et al. [1995,
1996, 1997] involved decoupling of the subducting crust
from the mantle lithosphere and exhumation of the coherent
nappe by faulting of the crust. Although overall kinematics
of exhumation resembles the eduction model, the dynamics
of these models are different. In Chemenda et al. [1995], the
buried crust decouples from the mantle lithosphere, implying
weak mechanical coupling at the Moho. Similarly to our
model, the exhumation of the nappe is coherent. The exhumation process differ by the fact that it requires the presence
13 of 17
B08411
B08411
DURETZ ET AL.: SLAB EDUCTION
of a thrust on the lower plate which is not required by the
eduction model. A main similarity between those models is
the presence of a normal sense shear zone in the vicinity of
the suture. In our simulations, the main normal shear zone is
produced consequently to slab detachment and eduction.
Despite slab breakoff occurred in the experiments of Boutelier
et al. [2004], no eduction was observed because of two major
reasons: (1) the crust had already delaminated from the mantle
at breakoff time, preventing coherent lithospheric-scale
motion, (2) the system was subject to a continued kinematic
push.
7.4. Dimensionality and Plate Motions
[39] Our two-dimensional models do not take into account
effects related to three dimensional flow in the mantle. We
can expect that the toroidal flow component can have a
strong influence on subduction-collision systems. A consequence is the tendency for subduction to roll back, leading to
a more decoupled, or retreating, style of collision [Husson
et al., 2009; Stegman et al., 2006]. The plate eduction endmember is triggered by slab detachment, recent threedimensional studies of slab detachment [Burkett and Billen,
2011; van Hunen and Allen, 2011] show that the tearing of
the slab can happen in an inhomogeneous manner along the
trench direction. We thus expect that plate eduction will take
place progressively in the along trench direction following
the detachment of the slab at depth. This direction of slab
tearing should also control the timing and location of
emplacement of the exhumed high pressure nappes as well
as surface uplift and filling of the adjacent basins. Another
3D aspect of the force balance change related to slab
detachment is the possibility to trigger abrupt changes in
plate motion [Austermann et al., 2011]. In our 2D study, the
main consequence of the slab pull is the partial eduction of
the subducted slab from the subduction zone. Our results
shows that a lower plate horizontal displacement on the
order of 100 km can result from eduction. In three dimensions, finite-width slabs associated with progressive (along
trench) slab detachment may lead to changes in plate motion
directions and potentially to rotations. On the other hand, we
could also expect that plate rotations, which can be triggered
by oblique continental collision [Bellhasen et al., 2003] can
strongly affect slab detachment [van Hunen and Allen, 2011]
and potentially help and/or trigger plate eduction in some
specific conditions. Another simplification of our 2D results
relies on the fact that detachment results in a total loss of the
slab pull force. In three-dimensions, we may expect that slab
portions might remain attached in the along-trench direction
(along strike coupling) and provide additional pull force.
The amount of eduction will consequently decrease according to the magnitude of pull that can be transmitted along
strike. Variations of slab width may also occur during subduction [Guillaume et al., 2010] and might strongly perturb
the subduction force leading to variations of plate velocities,
such effect may promote slab rollback and detachment as
well as the development of slab tear faults [Wortel et al.,
2009].
8. Conclusions
[40] The plate eduction model is characterized by the
inversion of the subduction plane within continental collision
zones. This mechanism is likely to take place after slab
detachment has occurred and eliminated the effect of slab
pull. Our two-dimensional study demonstrates that this
model can partly explain the exhumation of buried continental crust and orogenic extension. Models that isolate the
plate eduction mechanism lead to quasi adiabatic decompression of the buried crust (≈2 GPa) in a timespan of 5 My.
The coherent extraction of the slab from the subduction zone
results in localized extensional strain within the subduction
channel. This response to slab detachment is accompanied by
the flattening of the slab and the build up of topography on
the lower plate. We have derived scaling laws that enable the
prediction of the lower plate velocity that is subjected to
eduction. The plate velocity is function of the buoyancy of
the previously subducted continental crust and of the lateral
force that is exerted on the collision zone (e.g. ridge push).
Asthenosphere and subduction channel viscosity mantle
have a first order influence on the rate of eduction and allow
eduction if their respective magnitudes are lower than 1022
and 1021 Pa.s. Eduction subsequent to slab detachment can
occur in combination to other geodynamic processes (e.g.
slab retreat) and may play a significant role in the geodynamic evolution of collision zones.
Appendix A: Numerical Code Description
[41] The thermomechanical code I2VIS solves the twodimensional steady state Stokes equations and heat conservation equation using the finite difference/marker-in-cell
method [Gerya and Yuen, 2003a; Gerya, 2010]:
∂sij
¼ #rgi
∂xj
ðA1Þ
∂vi
¼0
∂xi
ðA2Þ
!
"
∂
∂T
DT
#H
¼ #rCp
k
∂xi
∂xi
Dt
ðA3Þ
where xj represents the spatial coordinates, r, the material
density (kg/m3), k, the thermal conductivity (W/m/K), Cp,
the specific heat capacity (J/kg) and H (J/m3/s), the contribution of the different heat sources (radiogenic, shear, and
adiabatic heating). The density and heat capacity of each
lithology are functions of both pressure P (Pa) and temperature T (K). The oceanic crust, lithospheric and asthenospheric
densities are pre-computed via Gibbs free energy minimization and updated at each timestep. The densities of continental crust rocks are calculated via the equation of state:
#
#
#
$$
r ¼ r0 1 # bðT # 298:15Þ 1 þ a P ! 10#8 # 10#3
ðA4Þ
where r0 corresponds to the reference density (2700 kg/m3
for the upper crust and 2800 kg/m3 for the lower crust), b to
the isothermal compressibility (0.5 1/K for the crust) and a
to the thermal expansivity (1.5 1/kbar for the crust). The
thermal conductivity k depends on the temperature [Clauser
and Huenges, 1995], the functions used to evaluate k are
listed in Table 1.
14 of 17
B08411
[42] The mechanical solver uses a viscous formulation and
the deviatoric stress tensor sij relates to the material viscosity h and the rate of deformation !_ ij via:
sij ¼ 2h_! ij ¼ heff
!
"
∂vi ∂vj
þ
∂xj ∂xi
1
n
1#n
2n
hcreep ¼ h0 !_ II
!
"
Ea þ PVa
exp
nRT
ðA6Þ
where n is the stress exponent, h0, the reference viscosity
(Pan.s), Ea, the activation energy (J), Va, the activation volume (J/bar) and R, the gas constant (8.314472 J/mol/K).
[43] Mohr-Coulomb (or Drucker-Prager) plasticity act as a
stress limiter in the regions where the second stress invariant
(sII) exceeds the material yield stress. The yield stress
depends on the pressure, the cohesion C (MPa), and the
internal friction angle f. The stress is limited via local viscosity reductions such as
hcreep ≤
C þ P sinðfÞ
pffiffiffiffiffi
2 !_ II
ðA7Þ
In the mantle, at sufficiently high stress and low temperature,
Peierls plasticity may be the dominant deformation mechanism [Evans and Goetze, 1979; Kameyama et al., 1999;
Raterron et al., 2004; Katayama and Karato, 2008]. This
regime has exponential dependance on the second stress
invariant and can therefore act as a strong weakening
mechanism in the lithospheric mantle. The effective viscosity
corresponding to the Peierls creep regime is formulated as:
1
APeierls sII
"!
exp
Ea # PVa
RT
"
1#
!
sII
sPeierls
"k !q #
ðA8Þ
In our simulations, this mechanism becomes active when the
Peierls viscosity hPeierls is inferior to the creep viscosity hcreep.
We use the dry olivine parameters APeierls = 107.8 ! 10#12
and sPeierls = 9.1 GPa [Evans and Goetze, 1979]. The viscosity is limited such that 1018 < h < 1025 Pa.s.
[44] The model’s surface h (air/crust interface) evolves
following a gross-scale erosion-sedimentation law [Gerya
and Yuen, 2003b; Gerya, 2010]:
∂h
∂h
¼ vy # vx # e_ þ s_
∂t
∂x
[45] The result of the calculation of the fluid velocity
obtained on the Eulerian grid and is interpolated to the
Lagrangian markers, the advection equation is the solved
explicitly by a coordinate update of the Lagrangian markers:
ðA5Þ
The different lithologies deform according to a visco-plastic
rheology. At stresses larger than 30 kPa, most of the flow
occurs in the dislocation creep regime [Turcotte and
Schubert, 1982] and depends on the second invariant of
the strain rate tensor (_! II ), the temperature and the pressure
[Ranalli, 1995]. The effective viscosity corresponding to a
dislocation creep regime is calculated as following:
hPeierls ¼
B08411
DURETZ ET AL.: SLAB EDUCTION
ðA9Þ
where vy and vx are the uplift and advection velocity (m/s)
predicted by the tectonic model. ė and s_ represent prescribed
erosion and sedimentation rates (m/s), they are set to 0.1 mm/y
in our reference run. Erosion is active above a reference altitude of 1 km whereas sedimentation takes place at depths
below #1 km.
xtþ1
¼ xtj þ Dt vt
j
ðA10Þ
where vt corresponds to the marker velocity computed vi 4th
order (in space) Runge-Kutta scheme, Dt represent the
value of the timestep at the current time integration.
[46] Acknowledgments. T.D. was funded by the SNF-EU research
grant 20TO21-120535 (TOPO-4D). A Centre of Excellence grant from
the Norwegian Research Council to PGP financed T.B. Andersen’s work
in this study. Boris Kaus was supported by starting grant 258830 by the
European Research Council. Careful reviews from D. Boutelier, E Burkett,
and W. P. Schellart have helped to improve the original manuscript. We
thank Jean-Pierre Burg for proofreading the manuscript and discussions.
This work benefited from helpful discussions with L. Le Pourhiet and
E. Burov.
References
Andersen, T. B. (1998), Extensional tectonics in the Caledonides of southern
Norway: An overview, Tectonophysics, 285, 333–351.
Andersen, T. B., and B. Jamtveit (1990), Uplift of deep crust during
orogenic extensional collapse: A model based on field studies in the
Sogn-Sunnfjord region of Western Norway, Tectonics, 9(4), 1097–1111.
Andersen, T. B., K. P. Skjerlie, and H. Furnes (1990), The Sunnfjord
Melange, evidence of Silurian ophiolite accretion in the West Norwegian
Caledonides, J. Geol. Soc., 147, 59–68.
Andersen, T. B., B. Jamtveit, J. F. Dewey, and E. Swensson (1991),
Subduction and eduction of continental crust: Major mechanisms during
continent-continent collision and orogenic extensional collapse, a model
based on the south Norwegian Caledonides, Terra Nova, 3, 303–310.
Andrews, E. R., and M. I. Billen (2009), Rheologic controls on the dynamics
of slab detachment, Tectonophysics, 464(1–4), 60–69.
Austermann, J., Z. Ben-Avraham, P. Bird, O. Heidbach, G. Schubert, and
J. M. Stock (2011), Quantifying the forces needed for the rapid change
of Pacific plate motion at 6 Ma, Earth Planet. Sci. Lett., 307, 289–297.
Austrheim, H., F. Corfu, I. Bryhni, and T. B. Andersen (2003), The Proterozoic Hustad igneous complex: A low strain enclave with a key to the
history of the Western Gneiss Region of Norway, Precambrian Res.,
120, 149–175.
Baumann, C., T. V. Gerya, and J. A. D. Connolly (2009), Numerical
modelling of sponaneous slab breakoff dynamics during continental collision, in Advances in Interpretation of Geological Processes: Refinement
of Multi-scale Data and Integration in Numerical Modelling, edited by
M. Spalla, A. Marotta, and G. Gosso, Geol. Soc. Spec. Publ., 322, 99–114.
Beaumont, C., R. A. Jamieson, M. H. Nguyen, and B. Lee (2001), Himalayan
tectonics explained by extrusion of a low-viscosity crustal channel coupled
to focused surface denudation, Nature, 414(6865), 738–742.
Beaumont, C., R. A. Jamieson, M. H. Nguyen, and S. Medvedev (2004),
Crustal channel flows: 1. Numerical models with implications to the tectonics of the Himalayan-Tibetan orogen, J. Geophys. Res., 109, B06406,
doi:10.1029/2003JB002809.
Beaumont, C., R. A. Jamieson, M. H. Nguyen, and S. Ellis (2006), Crustal
flow modes in large hot orogens, in Channel Flow, Ductile Extrusion
and Exhumation in Continental Collision Zones, edited by R. D. Law,
M. P. Searle, and L. Godin, Geol. Soc. Spec. Publ., 268, 91–145.
Bellhasen, N., C. Faccenna, F. Funiciello, J. M. Daniel, and L. Jolivet
(2003), Why did Arabia separate from Africa? Insights from 3-D laboratory experiments, Earth Planet. Sci. Lett., 216, 365–381.
Bialas, R. W., F. Funiciello, and C. Faccenna (2011), Subduction and exhumation of continental crust: Insights from laboratory models, Geophys.
J. Int., 184, 43–64.
Bird, P. (1978), Initiation of intracontinental subduction in the Himalaya,
J. Geophys. Res., 83(B10), 4975–4987.
Boutelier, D., A. Chemenda, and C. Jorand (2004), Continental subduction
and exhumation of high-pressure rocks: Insights from thermo-mechanical
laboratory modelling, Earth Planet. Sci. Lett., 222, 209–216.
Brueckner, H. K. (2009), Subduction of continental crust, the origin of postorogenic granitoids (and anorthosites?) and the evolution of Fennoscandia,
J. Geol. Soc., 166, 753–762.
Brueckner, H. K., and H. L. M. van Roermund (2004), Dunk tectonics:
A multiple subduction/eduction model for the evolution of the Scandinavian
Caledonides, Tectonics, 23, TC2004, doi:10.1029/2003TC001502.
15 of 17
B08411
DURETZ ET AL.: SLAB EDUCTION
Brun, J. P., and C. Facenna (2008), Exhumation of high-pressure rocks
driven by slab rollback, Earth Planet. Sci. Lett., 272, 1–7.
Buiter, S. J. H., R. Govers, and M. J. R. Wortel (2002), Two-dimensional
simulation of surface deformation caused by slab detachment, Tectonophysics, 354(3–4), 192–210.
Buffet, B. A. (2006), Plate force due to bending at subduction zones,
J. Geophys. Res., 111, B09405, doi:10.1029/2006JB004295.
Burkett, E. R., and M. I. Billen (2011), Three-dimensionality of slab detachment due to ridge-trench collision: Laterally simultaneous boudinage
versus tear propagation, Geochem. Geophys. Geosyst., 11, Q11012,
doi:10.1029/2010GC003286.
Chemenda, A. I., M. Mattauer, J. Malavieille, and A. Bokun (1995),
A mechanism for syn-collisional rock exhumation and associated normal
faulting: Results from physical modelling, Earth Planet. Sci. Lett., 132,
225–232.
Chemenda, A. I., M. Mattauer, and A. Bokun (1996), Continental subduction and a mechanism for exhumation of high-pressure metamorphic
rocks: New modelling and field data from Oman, Earth Planet. Sci. Lett.,
143, 172–182.
Chemenda, A. I., P. Matte, and V. Sokolov (1997), A model of Paleozoic
obduction and exhumation of high-pressure/low-temperature rocks in
the southern Urals, Tectonophysics, 276, 217–227.
Clauser, C., and E. Huenges (1995), Thermal conductivity of rocks and
minerals, in Rock Physics and Phase Relations, AGU Ref. Shelf vol. 3,
edited by T. J. Ahrens, pp. 105–126, AGU, Washington, D. C.
Cloos, M., and R. L. Shreve (1988), Subduction-channel model of prism
accretion, melange formation, sediment subduction, and subduction erosion at convergent plate margins: 1. Background and description, Pure
Appl. Geophys., 128(3–4), 455–500.
Corfu, F., T. H. Torsvik, T. B. Andersen, L. D. Ashwal, D. M. Ramsay, and
R. J. Roberts (2006), Early Silurian mafic-ultramafic and granitic plutonism
in contemporaneous flysch, Magerøy, northern Norway: U-Pb ages and
regional significance, J. Geol. Soc., 163, 291–301.
Crameri, F., H. Schmeling, G. J. Golabek, T. Duretz, R. Orendt,
S. J. H. Buiter, D. A. May, B. J. P. Kaus, T. V. Gerya, and P. J. Tackley
(2012), A comparison of numerical surface topography calculations in geodynamic modelling: An evaluation of the ‘sticky air’ method, Geophys. J.
Int., 189, 38–54.
Davies, H. J., and F. von Blanckenburg (1995), Slab breakoff: A model of
lithosphere detachment and its test in the magmatism and deformation of
collisional orogens, Earth Planet. Sci. Lett., 129(1–4), 85–102.
Dixon, J. M., and E. Farrar (1980), Ridge subduction, eduction, and the
neogene tectonics of southwestern North America, Tectonophysics, 67,
81–99.
Duretz, T., T. V. Gerya, and D. A. May (2011), Numerical modelling of
spontaneous slab breakoff and subsequent topographic response, Tectonophysics, 502(1–2), 244–256, doi:10.1016/j.tecto.2010.05.024.
Dvorkin, J., A. Nur, G. Mavko, and Z. Ben-Avraham (1993), Narrow
subducting slabs and the origin of backarc basins, Tectonophysics,
227, 63–73.
Evans, B., and C. Goetze (1979), Temperature variation of hardness of
olivine and its implication for polycristalline yield stress, J. Geophys.
Res., 84, 5505–5524.
Fossen, H. (2000), Extensional tectonics in the Caledonides: Synorogenic
or postorogenic?, Tectonics, 19(2), 213–224.
Froitzheim, N., J. Pleuger, S. Roller, and T. Nagel (2003), Exhumation of
high- and ultrahigh-pressure metamorphic rocks by slab extraction,
Geology, 31, 925–928.
Gerya, T. V. (2010), Introduction to Numerical Geodynamic Modelling,
345 pp., Cambridge Univ. Press, Cambridge, U. K.
Gerya, T. V., and D. A. Yuen (2003a), Charaterictics-based marker method
with conservative finite-difference schemes for modeling geological
flows with strongly variable transport properties, Phys. Earth Planet.
Inter., 140(4), 293–318.
Gerya, T. V., and D. A. Yuen (2003b), Rayleigh-Taylor instabilities from
hydration and melting propel ‘cold plumes’ at subduction zones, Earth
Planet. Sci. Lett., 212(1–2), 47–62.
Gerya, T. V., B. Stoeckhert, and L. L. Perchuk (2002), Exhumation of highpressure metamorphic rocks in a subduction channel: A numerical
simula- tion, Tectonics, 21(6), 1056, doi:10.1029/2002TC001406.
Gerya, T. V., D. A. Yuen, and W. Maresch (2004), Thermomechanical
modelling of slab detachment, Earth Planet. Sci. Lett., 226(1–2),
101–116.
Glodny, J., A. Kühn, and H. Austrheim (2008), Geochronology of
fluid-induced eclogite and amphibolite facies metamorphic reactions in a
subduction-collision system, Bergen Arcs, Norway, Contrib. Mineral.
Petrol., 156, 27–48.
B08411
Guillaume, B., F. Funiciello, C. Faccenna, J. Martinod, and V. Olivetti
(2010), Spreading pulses of the Tyrrhenian Sea during the narrowing of
the Calabrian slab, Geology, 38(9), 819–822.
Hacker, B. R., T. B. Andersen, S. Johnston, A. R. C. Kylander-Clark,
E. M. Peterman, E. O. Walsh, and D. Young (2010), High-temperature
deformation during continental-margin subduction and exhumation: The
ultrahigh-pressure Western Gneiss Region of Norway, Tectonophysics,
480(1–4), 149–171.
Huet, B., L. Le Pourhiet, L. Labrousse, E. Burov, and L. Jolivet (2010),
Post-orogenic extension and metamorphic core complexes in a heterogeneous crust: The role of crustal layering inherited from collision: Application to the Cyclades (Aegean domain), Geophys. J. Int., 184(2), 611–625.
Husson, L., J.-P. Brun, P. Yamato, and C. Faccenna (2009), Episodic slab
rollback fosters exhumation of HP–UHP rocks, Geophys. J. Int., 179,
1292–1300.
Janák, M., N. Froitzheim, M. Vrabec, E. J. Krogh Ravna, and J. C. M. De Hoog
(2006), Ultrahigh-pressure metamorphism and exhumation of garnet
peridotite in Pohorje, Eastern Alps, J. Metamorph. Geol., 24(1), 19–31,
doi:10.1111/j.1525-1314.2005.00619.x.
Kameyama, M., D. A. Yuen, and S.-I. Karato (1999), Thermal-mechanical
effects of low-temperature plasticity (the Peierls mechanism) on the
deformation of viscoelastic shear zone, Earth Planet. Sci. Lett., 168(1–2),
159–172.
Katayama, I., and S.-I. Karato (2008), Rheological structure and deformation of subducted slabs in the mantle transition zone: Implications
for mantle circulation and deep earthquakes, Phys. Earth Planet. Inter.,
168(3–4), 125–133.
Kaus, B. (2010), Factors that control the angle of shear bands in geodynamic
numerical models of brittle deformation, Tectonophysics, 484(1–4), 36–47.
Krabbendam, M., A. Wain, and T. B. Andersen (2000), Pre-Caledonian
granulite and gabbro enclaves in the Western Gneiss Region, Norway:
Indications of incomplete transition at high pressure, Geol. Mag., 137,
235–255.
Krogh, T. E., S. L. Kamo, P. Robinson, M. P. Terry, and K. Kwok (2011),
U-Pb zircon geochronology of eclogites from the Scandian Orogen,
northern Western Gneiss Region, Norway: 14–20 million years between
eclogite crystallization and return to amphibolite- facies conditions,
Can. J. Earth Sci., 48, 441–472.
Labrousse, L., L. Jolivet, T. B. Andersen, P. Agard, H. Maluski, and
U. Schärer (2004), Pressure-temperature-time deformation history of the
exhumation of ultra-high pressure rocks in the Western Gneiss Region,
Norway, in Gneiss Domes in Orogeny, vol. 380, edited by D. L. Whitney,
C. Teyssier, and C. S. Siddoway, pp. 155–185, Geol. Soc. Am., Boulder,
Colo.
Labrousse, L., G. Hetény, H. Raimbourg, L. Jolivet, and T. B. Andersen
(2010), Initiation of crustal-scale thrusts triggered by metamorphic
reactions at depth: Insights from a comparison between the Himalayas
and Scandinavian Caledonides, Tectonics, 29, TC5002, doi:10.1029/
2009TC002602.
Lallemand, S., A. Heuret, C. Faccenna, and F. Funiciello (2008), Subduction dynamics as revealed by trench migration, Tectonics, 27, TC3014,
doi:10.1029/2007TC002212.
Li, Z., and T. V. Gerya (2009), Polyphase formation and exhumation of
high- to ultrahigh-pressure rocks in continental subduction zone: Numerical
modeling and application to the sulu ultrahigh-pressure terrane in eastern
China, J. Geophys. Res., 114, B09406, doi:10.1029/2008JB005935.
Li, Z. H., Z. Q. Xu, and T. V. Gerya (2011), Flat versus steep subduction:
Contrasting modes for the formation and exhumation of high- to ultrahighpressure rocks in continental collision zones, Earth Planet. Sci. Lett., 301,
65–77.
Lister, G., M. A. Forster, and T. J. Rawling (2001), Episodicity during
orogenesis, Geol. Soc. Spec. Publ., 184, 89–113.
Manea, V., M. Manea, V. Kostoglodow, and G. Sewell (2006), Intraslab
seismicity and thermal stress in the subducted Cocos plate beneath central
Mexico, Tectonophysics, 420, 389–408.
McKenzie D. P. (1969), Speculations on the consequences and causes of
plate motions, Geophys. J. R. Astron. Soc., 18, 1–32.
Mishin, Y. A., T. V. Gerya, J.-P. Burg, and J. A. Connolly (2008), Dynamics
of double subduction: Numerical modeling, Phys. Earth Planet. Inter.,
171(1–4), 280–295, doi:10.1016/j.pepi.2008.06.012.
Ranalli, G. (1995), Rheology of the Earth, 2nd ed., Chapman and Hall,
London.
Raterron, P., Y. Wu, D. J. Weidner, and J. Chen (2004), Low-temperature
olivine rheology at high pressure, Phys. Earth Planet. Inter., 145, 149–159.
Rey, P., J. P. Burg, and M. Casey (1997), The Scandinavian Caledonides
and their relationship to the Variscan belt, Geol. Soc. Spec. Publ., 121,
179–200.
Røhr, T. S., F. Corfu, H. Austrheim, and T. B. Andersen (2004),
Sveconorwegian U-Pb zircon and monazite ages of granulite-facies
16 of 17
B08411
DURETZ ET AL.: SLAB EDUCTION
rocks, Hisarøya Gulen, Western Gneiss Region, Norweg. J. Geol., 84,
251–256.
Schellart, W. P., G. Lister, and V. G. Toy (2006), A Late Cretaceous and
Cenozoic reconstruction of the Southwest Pacific region: Tectonics controlled by subduction and slab rollback processes, Earth Sci. Rev., 76,
191–233.
Schlindwein, V., and W. Jokat (2000), Post-collisional extension of the East
Greenland Caledonides: A geophysical perspective, Geophys. J. Int., 140,
559–567.
Schmalholz, S. M. (2011), A simple analytical solution for slab detachment,
Earth Planet. Sci. Lett., 304(1–2), 45–54.
Schmalholz, S. M., D. W. Schmid, and R. C. Fletcher (2008), Evolution of
pinch-and-swell structures in a power-law layer, J. Struct. Geol., 30(5),
649–663.
Schmeling, H., et al. (2008), A benchmark comparison of spontaneous
subduction models: Towards a free surface, Phys. Earth Planet. Inter.,
171(1–4), 198–223.
Schneider, D. A., S. J. Zahniser, J. M. Glascock, S. M. Gordon, and
M. Manecki (2006), Thermochronology of the West Sudetes (Bohemian
Massif): Rapid and repeated eduction in the Eastern Variscides, Poland
and Czech Republic, Am. J. Sci., 306, 846–873.
Scholz, C. H., and J. Campos (1995), On the mechanism of seismic
decoupling and back arc spreading at subduction zones, J. Geophys.
Res., 100(B11), 103–115.
Stegman, D. R., J. Freeman, W. P. Schellart, L. Moresi, and D. A. May
(2006), Influence of trench width on subduction hinge retreat rates in
3-D models of slab rollback, Geochem. Geophys. Geosyst., 7, Q03012,
doi:10.1029/2005GC001056.
Stevenson, D. J., and J. S. Turner (1977), Angle of subduction, Nature,
270(1–2), 334–336.
Stöckhert, B., and T. V. Gerya (2005), Pre-collisional high pressure metamorphism and nappe tectonics at active continental margins: A numerical
simulation, Terra Nova, 17, 102–110.
Tucker, R. D., P. Robinson, A. Solli, D. G. Gee, T. Thorsnes, T. E. Krogh,
Ø. Nordgulen, and M. E. Bickford (2004), Thrusting and extension in the
Scandian Hinterland, Norway: New U-Pb ages and tectonostratigraphic
evidence, Am. J. Sci., 304, 477–532.
B08411
Turcotte, D. L., and G. Schubert (1982), Geodynamics: Application of
Continuum Physics to Geological Problems, John Wiley, New York.
van Hunen, J., and M. B. Allen (2011), Continental collision and slab
break-off: A comparison of 3-D numerical models with observations,
Earth Planet. Sci. Lett., 302(1–2), 27–37.
van Roermund, H. (2009), Recent progress in Scandian ultrahigh-pressure
metamorphism in the northernmost domain of the Western Gneiss
Complex, SW Norway: Continental subduction down to 180–200 km
depth, Am. J. Sci., 166(4), 739–751.
Vanderhaeghe, O., and C. Teyssier (2001), Partial melting and flow of
orogens, Tectonophysics, 342(3–4), 451–472.
Vrijmoed, J. C., Y. Y. Podladchikov, T. B. Andersen, and E. H. Hartz
(2009), An alternative model for ultra-high pressure in the Svartberget
Fe-Ti garnet-peridotite, Western Gneiss Region, Norway, Eur. J. Mineral.,
21(6), 1119–1133.
Warren, C. J., C. Beaumont, and R. A. Jamieson (2008a), Modelling
tectonic styles and ultra-high pressure (UHP) rock exhumation during the
transition from oceanic subduction to continental collision, Earth Planet.
Sci. Lett., 267, 129–245.
Warren, C. J., C. Beaumont, and R. A. Jamieson (2008b), Formation and
exhumation of ultra-high-pressure rocks during continental collision: Role
of detachment in the subduction channel, Geochem. Geophys. Geosyst., 9,
Q04019, doi:10.1029/2007GC001839.
Wortel, R., R. Govers, and W. Spakman (2009), Continental collision and
the STEP-wise evolution of convergent plate boundaries: From structure
to dynamics, in Subduction Zones Geodynamics, Front. Earth Sci., edited
by S. Lallemand and F. Funiciello, pp. 47–59, Springer, Berlin.
Yamato, P., E. Burov, P. Agard, L. Le Pourhiet, and L. Jolivet (2008),
HP-UHP exhumation during slow continental subduction: Self-consistent
thermodynamically and thermomechanically coupled model with application to the Western Alps, Earth Planet. Sci. Lett., 271(1–4), 63–74.
Young, D. J., B. R. Hacker, T. B. Andersen, and F. Corfu (2007), Prograde
amphibolite facies to ultrahigh-pressure transition along Nordfjord, western
Norway: Implications for exhumation tectonics, Tectonics, 26, TC1007,
doi:10.1029/2004TC001781.
17 of 17
Download